Вы находитесь на странице: 1из 8

Full Paper

Tacticity as a Factor Contributing to the Thermal Stability of Polystyrene


Kai Chen, Kehley Harris, Sergey Vyazovkin*

TGA has been used to study the thermal degradation of PS in atactic (aPS), syndiotactic (sPS), and isotactic (iPS) forms. Thermal stability of sPS is similar to that of aPS, whereas iPS degrades at a notably higher temperature. TGA-FT-IR data do not show any obvious difference in the degradation products of the three materials. Isoconversional kinetic analysis of nonisothermal TGA data yields somewhat higher activation energy for degradation of iPS than for aPS that is found to result in a smaller rate constant for iPS. Another important factor contributing to slower degradation rate of iPS is found to be the reaction model that has been experimentally determined from isothermal TGA measurements. The enhanced thermal stability of iPS correlates with its hindered molecular mobility that appears to originate from intramolecular connement. The hindered mobility has been detected by comparing the glass transition dynamics in amorphous iPS and aPS samples. By using multi-frequency temperature modulated DSC (TOPEM1) it has been demonstrated that the activation energy of the glass transition is noticeably larger for iPS than for aPS that supports the hypothesis of the slowed down molecular mobility.

Introduction
Thermal properties of polymeric materials are largely determined by the dynamics of the polymer chains. The chain dynamics can be altered by various treatments such as crosslinking, plasticizing, lling, drawing, blending and others. The treatments may either ease or constrain the motion of the polymer chains, therefore, altering their conformation compared to that in the virgin non-treated
K. Chen, K. Harris, S. Vyazovkin Department of Chemistry, University of Alabama at Birmingham, 901 S. 14th Street, Birmingham, AL 35294, USA E-mail: vyazovkin@uab.edu
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

polymer. In the recent years, polymer nanocomposites have drawn a great deal of research attention. The uniqueness of these systems is that they can constrain the polymer chain motion on the nanometer scale, creating so-called nanoconnement. For instance, intercalating a polymer chain into 15 nm gallery of layered silicate clay prevents the chain from assuming its natural conformation of a coil, whose radius of gyration Rg is typically on the order of 520 nm. Molecular dynamics (MD) simulations[14] suggest that nanoconnement is also accomplished near the clay surface, where the chains organize themselves in layers, whose density alternates being respectively higher and lower than the bulk density. This nanoconnement has been linked to the coexistence

DOI: 10.1002/macp.200700426

2525

K. Chen, K. Harris, S. Vyazovkin

of fast and slow segmental dynamics revealed by MD simulations and NMR measurements.[3,4] Nanoconnement is also accomplished in polymer brushes, whose bristles are made of uncoiled polymer chains. Our recent studies[59] on a polystyrene (PS)-clay brush system indicate that average interchain distance in it is about half that found in regular PS,[8] so that individual chains experience strong nanoconnement from their neighbors. The resulting enhanced interchain interaction correlates with increases in the activation energy of the glass transition,[6,7] the temperature and activation energy of degradation,[5,6] and in the formation of interchain reaction product a-methylstyrene.[8,9] The aforementioned types of nanoconnement are primarily intermolecular in their nature as they force interaction of an individual chain with other chains and/or with a surface. However, it is possible to conne the intramolecular motion of individual chains by introducing, for example, tacticity. Tacticity is known to affect the polymer chain conformation and dynamics. MD[10] and Monte Carlo[11] simulations indicate that in the melt syndiotactic poly(propylene) (PP) adopts conformations, having larger square end-to-end distance and statistical segment length than those in isotactic or atactic PP. The computational predictions agree with small angle neutron scattering (SANS) measurements[12] and intrinsic viscosity data.[11] For poly(methyl methacrylate) (PMMA), SANS measurements[13] demonstrate that the isotactic form has a larger radius of gyration than the syndiotactic one. For PS, the solid state 13C NMR and IR spectroscopy indicate[14] that in the glassy state the syndiotactic form has a larger content of the trans conformations than the isotactic or atactic forms. The tacticity-induced conformational changes are accompanied by a change in the chain dynamics. MD simulations and NMR experiments demonstrate an increase in the rate of the segmental relaxation dynamics of PP in the row: syndiotactic, atactic, isotactic.[10,15] Rheological measurements show that syndiotactic PP exhibits a 10 times higher zero shear rate viscosity[16] and 1015 kJ mol1 larger activation energy of viscous ow[16,17] than isotactic PP. Similarly, a larger zero shear rate viscosity is found in syndiotactic PMMA.[1820] Undoubtedly, the segmental dynamics plays an important role in the polymer reactivity because it is closely related to the basic kinetic factors such as the frequency of collisions and the rate of mass transfer. For this reason, tacticity can be a factor affecting the thermal degradation and, therefore, thermal stability of polymers. Surprisingly, the literature reports very little on the effect of tacticity on the thermal degradation kinetic of polymers. In an inert atmosphere, the syndiotactic form of PMMA displays a higher degradation temperature than the isotactic one[21] that is attributed to the slower chain mobility of the
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

former. A study of the thermo-oxidative degradation of PP in air demonstrates that the syndiotactic form is more stable than isotactic one.[22] There have been a few thermal degradation studies of isotactic[23,24] (iPS) and syndiotactic[25] (sPS) PS. However, no degradation studies of all three tactic forms under comparable conditions (i.e., similar molecular weight, atmosphere, experimental setup) have been performed and no reliable kinetic parameters have been reported. Nevertheless, the rotational isomeric state model predicts[26] iPS and sPS to have a larger intrinsic viscosity than atactic PS (aPS) that is consistent with the melt viscosity data[27] obtained for iPS and aPS of the similar molecular weight. Also, simulations[28,29] suggest that within the same chain the interactions between nearby phenyl rings are stronger in iPS than in sPS so that the former experiences a larger rotational energy barrier and, thus, larger steric hindrance. Based on these data one can certainly expect an effect of tacticity on the thermal degradation of PS. Revealing such effect is the purpose of the present study that explores the kinetics of the thermal degradation and relaxation of PS of different tacticity. To minimize other effects, the present study has been performed on PS of similar molecular weights and under a similar set of conditions. It should be noted that tacticity can only introduce some limited connement within short chain segments so that the resulting conformation of the whole chain as well as its interaction with other chains are not as dramatically affected as in typical nanocomposites whose properties originate from the enhancement of intermolecular interactions. Therefore, alteration of the thermal stability due to tacticity is expected to be smaller than that due to intermolecular connement in nanocomposites. Nevertheless, exploring this phenomenon is of fundamental signicance for elucidating the effect of conned chain dynamics on the thermal stability of polymers.

Experimental Part
Materials
iPS (90% mass fraction of isotactic triads, Mw 400 000) and sPS (90% mass fraction of syndiotactic triads, Mw 300 000) were purchased from Scientic Polymer Products, Inc. A reference aPS sample (Mw 410 000, Mw =Mn % 1.06) was obtained from Polymer Laboratories Ltd. The polydispersity of the tactic PS samples was not reported by the supplier, although it can be expected to have a rather large value for iPS that is typically polymerized by Ziegler-Natta catalysts. The resulting increase in lower molecular weight fraction may, in principle, diminish to some extent the thermal stability of PS. Nevertheless, our preliminary data showed that the degradation temperatures of iPS are noticeably larger than those for three different aPS samples (Mw 100 000, Mw =Mn % 2.4 from Alfa Aesar, Mw 350 000, Mw =Mn % 2.1 from

2526

DOI: 10.1002/macp.200700426

Tacticity as a Factor Contributing to the Thermal Stability of Polystyrene

Aldrich and Mw 410 000, Mw =Mn % 1.06 from Polymer Laboratories Ltd). On the other hand, the three aPS samples barely showed any difference in their decomposition temperatures that suggests that the effect of polydispersity is inferior to that of tacticity. Therefore, even if iPS has a large polydispersity, the tacticity effect discussed in this paper can only be larger in iPS of a smaller polydispersity.

isoconversional methods provide an effective way of dealing with complex kinetics of thermally stimulated processes in polymeric materials. The underlying assumption of the methods is that the reaction model is invariant to temperature or heating rate. By virtue of this assumption, at a constant extent of conversion the reaction rate is only a function of the temperature:   d lnda=dt Ea dT 1 R a (2)

Characterization
The thermal degradation kinetics have been measured as the temperature- and time-dependent mass loss by using thermogravimetric analyzer (Mettler-Toledo TGA/SDTA851e). For nonisothermal measurements, samples of %5 mg have been placed in 40 mL aluminum pans and heated in the owing atmosphere of N2 at a ow rate of 70 ml min1, from 25 to 600 8C at the heating rates of 2.5, 5.0, 7.5, 10.0 and 12.5 K min1. The buoyancy effect in TGA has been accounted for by performing empty pan runs and subtracting the resulting data from the subsequent sample mass measurements. For isothermal measurements, the samples have been heated at the temperature 375 8C until complete degradation. The degradation products evolved during the TGA runs have been analyzed in-situ by IR Spectroscopy (Nicolet Nexus 470 FTIR) coupled with Mettler-Toledo TGA/SDTA851e. The heating rate was 10 K min1, and FT-IR spectra were collected at 4 cm1 resolution. The glass transition (relaxation) kinetics have been determined by using the recently invented[30] technique of multi-frequency temperature-modulated differential scanning calorimetry (TOPEM1 DSC 823e by Mettler Toledo). By overlaying a series of stochastic temperature pulses of different duration with a temperature ramp at a constant underlying rate, one single TOPEM1 run allows for determining the frequency dependence of the complex heat capacity that can be conveniently used to characterize the glass transition dynamics. To prepare glassy samples for DSC measurements, %5 mg of a material has been placed in 40 mL Al pan, heated to 250 8C (iPS) or 200 8C (aPS), and quenched into liquid nitrogen. The glass transition has been measured in a heating mode. The temperature program for all measurements was dened by superimposing the underlying heating rate of 0.5 K min1 and a series of stochastic temperature pulses of the 1 K amplitude and the time between the pulses ranging from 25 s to 60 s. DSC has been calibrated by using Indium and Zinc standards. Although Equation (2) takes its origin in the single step rate Equation (1), Ea is assumed constant only for a certain extent of conversion and the narrow temperature region related to this conversion. That is, the isoconversional methods analyze the degradation kinetics by using a set of Equation (1), each of which is associated with a certain a and has its respective value of Ea. The latter has a meaning of the effective activation energy, and its variation with a and/or T can provide valuable insights into the mechanism and kinetics of complex processes.[32] The present study makes use of an advanced isoconversional method developed by Vyazovkin.[33,34] The method has two key advantages over the popular methods of Flynn and Wall[35] and Ozawa.[36] Firstly, it can treat the kinetics, occurring under arbitrary variation in temperature, T(t) that allows for accounting of experimentally detected self-heating/cooling. For a series of n experiments conducted under different programs, Ti(t), the activation energy is estimated at any given a by nding Ea, which minimizes the function[33,34]
n X n X J Ea ; Ti ta i1 j6i

FEa

J Ea ; Tj ta

(3)

where Z   Ea dt exp RTi t taDa


ta

J Ea ; Ti ta 

(4)

Analysis of the Degradation Kinetics


Because of its complexity, polymer degradation kinetics can rarely be described by a single-step rate equation[31]   da E kT f a A exp f a dt RT (1)

Secondly, performing integration over small time segments [Equation (4)] eliminates a systematic error[34] found in the Flynn and Wall and Ozawa methods when Ea varies considerably with a. In Equation (4), a is varied from Da to 1Da with a step Da typically chosen to be 0.02. The integral, J in Equation (3), is computed by the trapezoid rule. The minimization procedure is repeated for each value of a to establish the dependence Ea on a. The pre-exponential factor has been evaluated by using so-called articial isokinetic relationship,[37] ln Aj c dEj (5)

that holds throughout the whole range of the degradation temperatures, T, and conversions, a. In Equation (1), t is the time, T is the temperature, R is the gas constant, A is the pre-exponential factor, E is the activation energy, k(T) is the rate constant, and f(a) is the reaction model. As discussed in a review paper,[32]
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

where c and d are constants. Twelve pairs of lnAj and Ej have been determined by the Coats-Redfern method[38] by using twelve different reaction models.[39] The parameters c and d have been evaluated by tting lnAj and Ej data to Equation (5). Then the isoconversional Ea values have been substituted for Ej in Equation (5) to calculate the corresponding lnAa values. The reaction model f(a) has been determined independently by running isothermal TGA measurements at the temperature

www.mcp-journal.de

2527

K. Chen, K. Harris, S. Vyazovkin

375 8C. At a constant degradation temperature, the shape of the mass loss curve is only controlled by f(a) because k(T) in Equation (1) remains constant throughout an isothermal measurement. The time-dependent mass loss is used to calculate the extent of conversion of degradation as a function of time by a mi mt mi mf (6)

where mi and mf are, respectively, the sample mass at the initial and nal time of measurement, and mt is the mass measured at different degradation times. The rst derivative of a versus t (da/ dt) plot has been used for curve-tting to calculate and compare the reaction model for PS in different tactic forms. Figure 1. TGA curves for degradation of aPS, sPS and iPS at a heating rate of 2.5 K min1 in nitrogen. Inset shows the rst derivative of TGA curves (DTG).

Analysis of the Glass Transition Dynamics


The glass transition is the process of relaxation from the nonequilibrium glassy state to the equilibrium liquid state. The rate of the process can be expressed by a rst-order kinetic equation[40] dj 1 j je dt t (7)

where j and je are respectively the non-equilibrium and equilibrium values of an order parameter such as enthalpy, and t is the relaxation time. For a narrow temperature region, the temperature dependence of t can be approximated by the Arrhenius equation   E (8) t A exp RT The activation energy in Equation (8) can be determined from the frequency-dependence of the glass transition temperature (Tg). The dependence is readily obtained by TOPEM1 DSC that measures the glass transition at different frequencies, f, as the heat capacity step, the midpoint of which can be used as an estimate of Tg. Since t and f are reciprocally related, Equation (8) can be used to determine the activation energy of the glass transition as follows E R d ln f 1 dTg (9)

of sPS and iPS are shifted to a higher temperature compared to aPS. The shift is the largest for iPS. From the rst derivative of TGA curves, the temperature of a maximum degradation rate for the sPS and iPS is respectively found to be 3 8C and 7 8C higher than that for aPS, indicating that iPS is the most thermally stable form of PS. The gas phase degradation products of PS samples of different tacticities have been compared by performing in-situ IR analysis. As shown in Figure 2, iPS and sPS samples display practically the same IR absorption pattern as their atactic counterpart. The major degradation product in all three cases is styrene detected by the absorption bands at 910, 989 cm1 ( C H bending) and 1 630 cm1 (C C stretch). Note that intermolecular connement of aPS in clay nanocomposites gives rise to an unusual degradation product a-methylstyrene that can be detected by 2 974 cm1 absorption band.[8,9] The band does not appear in the spectra of degradation products of iPS and sPS. Apparently, the chain tacticity does not alter

The TOPEM1 DSC technique has an advantage of evaluating the activation energy in a single experiment. This eliminates run-to-run experimental errors unavoidable in a series of single frequency runs in temperature modulated DSC as well as in a series of single heating rate runs in regular DSC, which can also be used for estimating the value of E.[41]

Results and Discussion


Degradation Figure 1 displays TGA curves of three PS samples degraded in the atmosphere of N2. It shows that the mass loss curves
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 2. Gas phase IR spectra of degradation products for aPS, sPS and iPS.

2528

DOI: 10.1002/macp.200700426

Tacticity as a Factor Contributing to the Thermal Stability of Polystyrene

the PS degradation pathway and the respective yield of degradation products. Since the degradation mechanism of PS is independent on its tacticity, our following study is focused on the effect of tacticity on the kinetics of PS degradation. The kinetics of the thermal degradation of PS in the three tactic forms have been further analyzed by using the advanced isoconversional method. The isoconversional activation energies of degradation are plotted in Figure 3. It shows that the effective activation energies (Ea) of all three PS samples are practically independent of the degradation conversion that suggests that the overall degradation process is limited by a single step.[32] Figure 3 also shows that aPS has an effective activation energy about 10 kJ mol1 lower than sPS and iPS, while the difference in Ea between iPS and sPS is insignicant. The values of Ea have been used to calculate the pre-exponential factor, another important parameter in Arrhenius law that controls the rate of the degradation process. As seen in Figure 4, lnAa is not practically dependent on the conversion, and aPS has the smallest value of lnAa among the three stereoisomers. The obtained values of Aa and Ea have been employed to determine the rate constant of degradation. Since both values are practically constant during the whole degradation process, the respective averages have been used to determine the k(T) values. The latter are presented in Figure 5 in the form of the Arrhenius plots. It is seen that of all three forms, iPS has the smallest rate constant throughout the whole temperature region of degradation. The result is obviously consistent with the TGA data (Figure 1) that suggest iPS being more thermally stable. It is noteworthy that Aa and Ea make opposing contributions to k(T) for the thermal degradation of iPS. Compared to aPS, the respective Aa value is larger for iPS that suggests that degradation of the latter should be faster. On the contrary, iPS demonstrates

Figure 4. Dependence of the pre-exponential factor (A) on the extent of conversion (a) for the non-isothermal degradation of aPS, sPS and iPS.

larger Ea so that its degradation should be slower than that of aPS. Since the combined effect of these two parameters results in k(T) being smaller for iPS than for aPS, we can conclude that an increase in the activation energy makes the dominant contribution to the enhanced thermal stability of iPS. As indicated by Equation (1), the rate of a degradation is also determined by the reaction model f(a) that has been independently evaluated from the time-dependence of mass loss in isothermal TGA measurements. Figure 6 displays that the degradation of iPS reaches a smaller extent of conversion than aPS at the same degradation time and temperature, indicating a slower degradation rate for iPS. This is in agreement with our dynamic TGA measurements (Figure 1) that demonstrate delayed degradation of iPS relative to aPS. The inset of Figure 6 further shows that both aPS and iPS display the maximum

Figure 3. Dependence of the effective activation energy (Ea) on the extent of conversion (a) for the nonisothermal degradation of aPS, sPS and iPS.
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 5. Comparison of rate constants [k(T)] for the nonisothermal degradation of aPS, sPS and iPS.

www.mcp-journal.de

2529

K. Chen, K. Harris, S. Vyazovkin

Figure 6. The time dependence of the extent of conversion (a) for isothermal degradation of aPS and iPS at 375 8C. Inset shows the plots of the rate versus conversion.

rate at middle conversion that is characteristic of autocatalytic processes, whose rate can be described as da kT am 1 an dt

(10)

By tting the experimental rate curves (inset of Figure 6) to Equation (10), the reaction models for aPS and iPS have been determined and compared in Figure 7. It is clearly seen that for iPS the values of f(a) are smaller than those for aPS throughout the whole region of degradation. That is, the decreased value of f(a) also contributes to slowing the degradation rate of iPS. Putting the above results together, we can conclude that the tacticity of iPS shows the degradation rate by decreasing both f(a) and k(T) or, in other words, by making respectively weaker the dependence of the rate on both conversion and temperature. The altered degradation kinetics of tactic PS demonstrated in our TGA measurements compliments the

previous literature reports[21,22,42,43] addressing tacticity as factor that can affect the degradation behavior of polymers. Since the reported stability of tactic PP has been investigated[22] in the reactive atmosphere of air, which is fundamentally different from the inert environment used in the present study, it seems more appropriate to compare our results with the results of degradation study of tactic PMMA performed in the inert atmosphere. The thermal stability of PMMA has been found[21] to depend on its tacticity, and chain mobility has been proposed as a key factor controlling the polymer degradation. In comparison with isotactic PMMA, syndiotactic PMMA has more entangled chains with a closer entanglement spacing (a smaller entanglement molecular weight).[18,44] The stronger intermolecular interactions, presumably associated with stronger carbonyl-carbonyl dipole interactions of ester group in syndiotactic PMMA,[45] impose a larger connement on the mobility of syndiotactic chains.[19] The resulting constrained chain mobility contributes to the higher thermal stability of syndiotactic PMMA.[21] It should be noted that in PMMA the syndiotactic form is more stable than the isotactic one, whereas in PS the situation is reverse (Figure 1). This, however, is not very surprising because in PS the isotactic form experiences larger intramolecular connement than the syndiotactic one. The origins of the conned molecular mobility in tactic PS can be understood from the results of conformational analysis.[28] It shows that in iPS there are strong interactions between side phenyl rings along the same chain. This steric hindrance (intramolecular effect) due to bulky side groups gives rise to a higher rotational energy barrier because the skeletal rotations of iPS require more consecutive monomer units to move cooperatively than in sPS.[28,29] Since the rotations around chain bonds are a part of the segmental mobility, an increased rotational cooperativity in iPS appears to be an important factor that causes intramolecular connement of the PS chain in a wide temperature region, covering both glassy and melt state. Relaxation The glass transition dynamics has been characterized by applying TOPEM1 multi-frequency DSC. sPS and iPS are semi-crystalline materials and aPS is an entirely amorphous material. A connement effect of crystalline phase on the mobility of the amorphous phase has been reported for iPS[46] and other semi-crystalline polymers.[47,48] In order to identify the effect of tacticity on the PS chain mobility, the effect of crystalline phase must be removed. Our glass transition study has focused on iPS, which, according to the thermal degradation data, demonstrates a larger effect of tacticity than sPS. Amorphous iPS has been analyzed in comparison to aPS.

Figure 7. The tted reaction models [f(a) am (1 a)n] for aPS and iPS.
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

2530

DOI: 10.1002/macp.200700426

Tacticity as a Factor Contributing to the Thermal Stability of Polystyrene

The activation energy (E) of the glass transition has been evaluated from the frequency dependence of Tg [Equation (9)] determined in TOPEM1 measurements. An example of a typical TOPEM1 measurement (on iPS sample) is given in Figure 8. The real part of the complex heat capacity is evaluated at the frequencies, f, from 10 to 25 mHz. It shows that the glass transition temperature increases with increasing the frequency. The frequency dependence of the glass transition temperature is plotted in Figure 9. The slope of lnf versus 1/Tg plots yields an activation energy for the glass transition. For iPS the E value is 504 7 kJ mol1, which is ca. 27 kJ mol1 larger than that for aPS (E 477 5 kJ mol1). The larger E for iPS indicates that a process of molecular motion in the isotactic amorphous state encounters a larger energetic barrier than in the atactic amorphous state, therefore, making the chain mobility more hindered. Although the obtained result demonstrates hindered iPS chain mobility in the transitional glass to liquid state (lower temperature region), we believe that this effect is relevant to the melt state (higher temperature region). This conjecture is supported by the reported viscosity data of PS which are clearly tacticity-dependent. The rotational isomeric state model predicts[26] that iPS and sPS have a larger intrinsic viscosity than aPS. The measured melt viscosity of iPS is higher than that of aPS when they have the same molecular weight.[27] By extrapolating the viscosity data[49] of sPS and aPS to the same temperature, we nd that sPS also displays a larger viscosity than aPS in the melt state. An increase in the viscosity can signicantly affect the kinetics of degradation that occurs in viscous reaction media,[50] because the effective time of a process, tef, is generally determined by the sum of the characteristic time of a chemical reaction, tr, and the characteristic time for reactants to diffuse in the viscous

Figure 9. Activation energy (E) of the glass transition for iPS determined from the frequency-dependence of Tg.

medium, t:[51] tef tr t (11)

In its turn, the characteristic time of diffusion is directly related to viscosity as follows,[52] t 4pa3 h kB T (12)

where h is the viscosity of the medium, a is the molecular radius, and kB is Boltzmanns constant. Equation (11) and (12) suggest that an increase in viscosity can slow down the overall process by increasing its effective time. A correlation between the conned chain mobility and enhanced thermal stability of iPS stresses the importance of molecular mobility in controlling the reactivity of a polymer. Our results combined with the results of an earlier study of stereoregular PMMA systems,[21] suggest that stereoisomers with lower chain mobility (e.g., isotactic PS and syndiotactic PMMA) should generally posses higher thermal stability and degrade slower.

Conclusion
The thermal degradation study of PS of different tacticity has demonstrated that iPS has measurably larger thermal stability than aPS. No evidence has been found to suggest that the tacticity of PS may affect its degradation pathway. The enhanced thermal stability of iPS reveals itself in the rate constant and reaction model, the respective values of which are smaller than those found for aPS. While reproducibly detectable, both effects are fairly small, which is not unexpected because the experimentally measured increase in the degradation temperature for iPS is also relatively small. The relaxation kinetics study has yielded an activation energy which is noticeably larger for iPS than

Figure 8. Frequency dependence of the heat capacity for iPS evaluated by TOPEM1.
Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.mcp-journal.de

2531

K. Chen, K. Harris, S. Vyazovkin

for aPS that is indicative of more conned molecular mobility in the former. The observed correlation between the conned intramolecular chain mobility and enhanced thermal stability is consistent with the literature report on degradation of tactic PMMA. It can be concluded that the tactic form that forces stronger intramolecular connement and slower segmental mobility should generally have larger thermal stability. It is noteworthy that the effect of intramolecular connement on degradation and relaxation of tactic polymers appears qualitatively similar, albeit quantitatively smaller, to the effect of intermolecular connement frequently found in degradation and relaxation studies of polymer nanocomposites.

Acknowledgements: We thank Mettler-Toledo Inc. for donating the TGA instrument and loaning multi-frequency TMDSC (TOPEM1). Received: August 7, 2007; Revised: August 27, 2007; Accepted: August 30, 2007; DOI: 10.1002/macp.200700426 Keywords: connement; degradation; isotactic; mobility; syndiotactic

[1] D. Schmidt, D. Shah, E. P. Giannelis, Curr. Opin. Solid State Mater. Sci. 2002, 6, 205. [2] E. Manias, V. Kuppa, Eur. Phys. J. E 2002, 8, 193. [3] E. Manias, V. Kuppa, D. K. Yang, D. B. Zax, Colloid Surf. A 2001, 187188, 509. [4] D. B. Zax, D. K. Yang, R. A. Santos, H. Hegemann, E. P. Giannelis, E. Manias, J. Chem. Phys. 2000, 112, 2945. [5] S. Vyazovkin, I. Dranca, X. W. Fan, R. Advincula, Macromol. Rapid Commun. 2004, 25, 498. [6] S. Vyazovkin, I. Dranca, X. W. Fan, R. Advincula, J. Phys. Chem. B 2004, 108, 11672. [7] S. Vyazovkin, I. Dranca, J. Phys. Chem. B 2004, 108, 11981. [8] K. Chen, M. A. Susner, S. Vyazovkin, Macromol. Rapid Commun. 2005, 26, 690. [9] K. Chen, S. Vyazovkin, Macromol. Chem. Phys. 2006, 207, 587. [10] S. J. Antoniadis, C. T. Samara, D. N. Theodorou, Macromolecules 1999, 32, 8635. [11] T. M. Madkour, A. Soldera, Eur. Polym. J. 2001, 37, 1105. [12] T. D. Jones, K. A. Chafn, F. S. Bates, B. K. Annis, E. W. Hagaman, M. H. Kim, G. D. Wignall, W. Fan, R. Waymouth, Macromolecules 2002, 35, 5061. [13] J. M. OReilly, D. M. Teegarden, G. D. Wignall, Macromolecules 1985, 18, 2747. [14] T. Nakaoki, M. Kobayashi, J. Mol. Struct. 2003, 655, 343. [15] V. Arrighi, D. Batt-Coutrot, C. H. Zhang, M. T. F. Telling, A. Triolo, J. Chem. Phys. 2003, 119, 1271. [16] A. Eckstein, C. Friedrich, A. Lobbrecht, R. Spitz, R. Mulhaupt, Acta Polym. 1997, 48, 41.

[17] A. Eckstein, J. Suhm, C. Friedrich, R. D. Maier, J. Sassmannshausen, M. Bochmann, R. Mulhaupt, Macromolecules 1998, 31, 1335. [18] K. Fuchs, C. Friedrich, J. Weese, Macromolecules 1996, 29, 5893. [19] S. Doulut, P. Demont, C. Lacabanne, Macromolecules 2000, 33, 3425. [20] A. Soldera, Y. Grohens, Polymer 2004, 45, 1307. [21] K. Hatada, T. Kitayama, N. Fujimoto, T. Nishiura, J. Macromol. Sci., Part A: Pure Appl. Chem. 1993, 30, 645. [22] T. Hatanaka, H. Mori, M. Terano, Polym. Degrad. Stab. 1999, 64, 313. [23] J. Boon, G. Challa, Makromol. Chem. 1965, 84, 25. [24] A. Nakajima, F. Hamada, T. Shimizu, Makromol. Chem. 1966, 90, 229. [25] G. P. Ravanetti, M. Zini, Thermochim. Acta 1992, 207, 53. [26] H. Z. Ma, L. X. Zhang, J. M. Xu, Acta Polym. Sin. 2001, 6, 769. [27] J. Boon, G. Challa, P. H. Hermans, Makromol. Chem. 1964, 74, 129. [28] S. Bruckner, G. Allegra, P. Corradini, Macromolecules 2002, 35, 3928. [29] G. Allegra, Macromol. Symp. 2004, 218, 89. [30] J. E. K. Schawe, T. Hutter, C. Heitz, I. Alig, D. Lellinger, Thermochim. Acta 2006, 446, 147. [31] J. H. Flynn, Encyclopedia of polymer science and engineering, Vol. Suppl., Wiley, New York 1989. [32] S. Vyazovkin, N. Sbirrazzuoli, Macromol. Rapid Commun. 2006, 27, 1515. [33] S. Vyazovkin, J. Comput. Chem. 1997, 18, 393. [34] S. Vyazovkin, J. Comput. Chem. 2001, 22, 178. [35] J. H. Flynn, L. A. Wall, J. Polym. Sci., Polym. Lett. Ed. 1966, 4, 323. [36] T. Ozawa, Bull. Chem. Soc. Jpn. 1965, 38, 1881. [37] S. Vyazovkin, Int. J. Chem. Kinet. 1996, 28, 95. [38] A. W. Coats, J. P. Redfern, Nature 1964, 201, 68. [39] S. Vyazovkin, Int. Rev. Phys. Chem. 2000, 19, 45. [40] G. R. Strobl, The physics of polymers: concepts for understanding their structures and behavior, Springer, Berlin, New York 1997. [41] C. T. Moynihan, A. J. Easteal, J. Wilder, J. Tucker, J. Phys. Chem. 1974, 78, 2673. [42] M. Lazzari, T. Kitayama, K. Hatada, O. Chiantore, Macromolecules 1998, 31, 8075. [43] H. Nakatani, S. Suzuki, T. Tanaka, M. Terano, Polymer 2005, 46, 12366. [44] W. Souheng, J. Polym. Sci., Part B: Polym. Phys. 1989, 27, 723. [45] O. N. Tretinnikov, K. Ohta, Macromolecules 2002, 35, 7343. [46] B. Natesan, H. Xu, B. S. Ince, P. Cebe, J. Polym. Sci., Part B: Polym. Phys. 2004, 42, 777. [47] S. Z. D. Cheng, M. Y. Cao, B. Wunderlich, Macromolecules 1986, 19, 1868. [48] C. Schick, A. Wurm, A. Mohammed, Thermochim. Acta 2003, 396, 119. [49] R. Pantani, A. Sorrentino, Polym. Bull. 2005, 54, 365. [50] S. Vyazovkin, N. Sbirrazzuoli, Macromol. Chem. Phys. 2000, 201, 199. [51] D. A. Frank-Kamenetskii, Diffusion and heat transfer in chemical kinetics, Plenum, New York 1969. [52] P. J. W. Debye, Polar molecules, Dover Publications, New York 1960.

2532

Macromol. Chem. Phys. 2007, 208, 25252532 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

DOI: 10.1002/macp.200700426

Вам также может понравиться